You are on page 1of 86

Relativistic

Quantum
Mechanics
MPAGS

Dr. Thomas Blake

October 2020
Contents

1 Special relativity and Lorentz invariance 1


1.1 Objectives of this course . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Natural units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Invariances in physical theories . . . . . . . . . . . . . . . . . . . . . . . . 3
1.5 Form invariance in classical mechanics . . . . . . . . . . . . . . . . . . . . 3
1.6 Galilei transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.7 Special relativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.8 The Lorentz transformation . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Examples of Lorentz invariance –


Maxwell’s equations and the Klein-Gordon equation 9
2.1 Lorentz covariance of Maxwell’s equations . . . . . . . . . . . . . . . . . . 10
2.2 Quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 The Scrödinger equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 The Klein-Gordon equation . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.5 The Feynman-Stückleberg interpretation . . . . . . . . . . . . . . . . . . . 16

3 Perturbation theory and particle scattering 17


3.1 Time dependent perturbation theory . . . . . . . . . . . . . . . . . . . . . 18

4 Coulomb scattering of charged spin-0 particles 24


4.1 Motion of a charged particle in an EM field . . . . . . . . . . . . . . . . . 25
4.2 Feynman rules for spin-0 particles . . . . . . . . . . . . . . . . . . . . . . . 28
4.3 Propagators and time-ordering . . . . . . . . . . . . . . . . . . . . . . . . . 30

5 Observable quantities 32
5.1 Box normalisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.2 Computing observable quantities from amplitudes . . . . . . . . . . . . . . 35
5.3 Decay rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.4 Cross sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.5 Mandelstam variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

i
PP2 Relativistic Quantum Mechanics

6 Calculating cross-sections for spin-0 particle scattering 40


6.1 Spinless scattering: e− µ− → e− µ− . . . . . . . . . . . . . . . . . . . . . . . 41
6.2 Spinless scattering: e− e− → e− e− . . . . . . . . . . . . . . . . . . . . . . . 43
6.3 Infrared divergences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
6.4 Spinless scattering: e+ e− → e+ e− . . . . . . . . . . . . . . . . . . . . . . . 45

7 The Dirac equation 47


7.1 The Dirac equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
7.2 Free particle solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

8 Spin and antiparticles 53


8.1 Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
8.2 Antiparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
8.3 Coupling spin- 21 particles to an EM field . . . . . . . . . . . . . . . . . . . 60
8.4 Gyromagnetic ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

9 Coulomb scattering of charged spin- 12 particles 63


9.1 Spin-1/2 scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
9.2 Interfering diagrams and relative signs . . . . . . . . . . . . . . . . . . . . 67
9.3 Spin-1/2 scattering: e− µ− → e− µ− . . . . . . . . . . . . . . . . . . . . . . 68

10 Trace techniques and spin sums 71


10.1 Trace techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
10.2 Spin-1/2 scattering: e− µ− → e− µ− . . . . . . . . . . . . . . . . . . . . . . 74
10.3 Spin-1/2 scattering: e− e+ → µ− µ+ . . . . . . . . . . . . . . . . . . . . . . 76
10.4 Photons and polarisation vectors . . . . . . . . . . . . . . . . . . . . . . . 78

T. Blake ii contact: thomas.blake@cern.ch


Lecture 1

Special relativity and


Lorentz invariance

1
PP2 Relativistic Quantum Mechanics

1.1 Objectives of this course


The objectives of this lecture course are to:

• get familiar with the notation of relativistic equations;

• understand how to use perturbation theory to derive transition amplitudes;

• understand the properties of the Dirac and Klein-Gordon equations;

• and understand how to perform calculations based on Feynman rules (spin sums/trace
techniques etc).

I am indebted to Michal Kreps, whose lectures this course is based on. Much of the content
of the course can also be found in chapters 3, 4, 5 and 6 of

• Quarks and Leptons: An Introductory Course in Modern Particle Physics


by Halzen and Martin.

1.2 Introduction
Why do we need Relativistic Quantum mechanics? In high energy physics we study
processes that involve “very small things travelling very fast”. For each aspect we have a
well tested theory:

“very small things” ⇒ quantum mechanics

“travelling very fast” ⇒ special relativity

The unification of general relativity (GR) and quantum mechanics (QM) is beyond the
scope of the MPAGS courses. There are numerous mathematical (quantum theories of
gravity are non-renormalizable) and conceptual issues (to do with locailty and unitarity)
with combining GR and QM. Attempts to unify GR and QM include string theory and
loop quantum gravity.

T. Blake 2 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

1.3 Natural units


In particle physics the physical constants

c = 2.99792458 × 108 m s−1 ,


~ = 1.054571726(47) × 10−34 J s ,

appear in many places. It is convenient to choose a system of units where ~ = c = 1, such


that mass, momentum (mc) and energy (mc2 ) have the same units. This allows us to be
lazy when writing down our equations and we’ll do this frequently throughout the course.
If we want to, we can convert back to SI units using dimensional analysis.

1.4 Invariances in physical theories


Invariances play an important role in physical theories. In 1915 Emmy Noether demon-
strated that every (differentiable) symmetry of the action of a physical system has a
corresponding conservation law.
Examples are:
invariance under translations ⇒ conservation of momentum

invariance under rotations ⇒ conservation of angular momentum

time invariance (same at t and t0 ) ⇒ conservation of energy

1.5 Form invariance in classical mechanics


We will start by looking at what is known as form-invariance i.e. that the equations
describing a physical system are the same for all observers.
law i = jk → i0 = j 0 k 0
observer A observer A0
In classical mechanics we know that equations are invariant under translations and rotations
of coordinate axis. However, these invariances are not obvious if we do not formulate our
equations in a suitable language. In classical mechanics we use vectors.
If we consider two coordinate systems rotated with respect to each other. A vector A ~
in system S will correspond to a vector A~ 0 in system S 0 ,

~ = (a1 , a2 , a3 ) , A
A ~ 0 = (a01 , a02 , a03 ) . (1.1)
~ and A
The vectors A ~ 0 correspond to the same physical object (they are the same vector)
and have the same magnitude (length) in each of the coordinate systems
~ 2 = (a1 )2 + (a2 )2 + (a3 )2 = (a01 )2 + (a02 )2 + (a03 )2 .
|A| (1.2)

T. Blake 3 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

~ 2 , i.e. the length of the vector, is an “invariant” quantity. We can write


The quantity |A|

~ = a1 ê1 + a2 ê2 + a3 ê3 = a01 ê0 + a02 ê0 + a03 ê0 ,


A (1.3)
1 2 3

where ê1 , ê2 , ê3 and ê01 , ê02 , ê03 are basis vectors describing the coordinate system of S and
S 0 , respectively. The coordinates are related through a linear transformation (a rotation)
3
X
i0
a = Rji aj . (1.4)
j=1

Lets apply this to Newtons law. In S

F~ = m~a or F i = mai , (1.5)

where the index i refers to a component of F~ . What form does this take in S 0 ?

F i = mai (1.6)
3
X 3
X
Rij F i = m Rij ai ,
i=1 i=1
0j 0j
F = ma

The form invariance comes from having objects that transform in the same way on the
left- and right-hand side of the equation.

T. Blake 4 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

1.6 Galilei transformations


In classical mechanics, if there are two systems of coordinates moving with a constant
relative velocity v with respect to each other, we make a Galilei transformation

x01 = x1 − vt
x02 = x2
(1.7)
x03 = x3
t0 =t

Newtons law’s are invariant under this type of transformation, but Maxwell’s equations
are not. Under a Galilei transformation,

dx0 dx
= −v ⇒ c0 = c − v (1.8)
dt0 dt
i.e. the speed of light is different for the different observers.

1.7 Special relativity


In 1905 Einstein postulated that:

• the laws of physics are invariant (the same) for all inertial systems (systems of
coordinates moving with uniform velocity with respect to each other);

• the speed of light in the vacuum is the same for all observers.

This requires a much more general transformation as it is a time dependent invariance.

1.8 The Lorentz transformation


The appropriate transformation to leave c unchanged is the Lorentz transformation. The
Lorentz transformation corresponding to a boost along x1 is
 0    
ct γ −βγ 0 0 ct
 x01   −βγ γ 0 0   1 
 02  =    x2  , (1.9)
 x   0 0 1 0  x 
x03 0 0 0 1 x3

where
v 1
β= , γ=p (1.10)
c 1 − β2

T. Blake 5 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Here we have included ct as another component x0 in a 4-vector

xµ = (x0 , x1 , x2 , x3 ) (1.11)

We conventionally use Greek indices to refer to components of a 4-vector (e.g. µ, ν etc)


and roman indices (e.g. i, j, k etc) to refer to components of a spatial 3-vector.
Written in this way, it is obvious that Einsteins postulates lead to a mixing of space
and time. In order to form invariant equations in special relativity we need to use 4-
vectors. In 3D cartesian space we had 3-vectors (~x). In 4D Minkowski space we have
space-time vectors (xµ ). The Lorentz transformation can be thought of like a (generalised)
rotation in the Minkowski space. They are part of a larger group of transformations, which
includes translations, called the Poincaré group. Writing equations in the form of 4-vectors
and balancing indices guarantees form invariance (or Lorentz covariance) under Lorentz
transformations. As the Lorentz transformation is linear, it can be written as
4
X
x0µ = Λµν xν (1.12)
ν=1

At this point we will introduce Einstein summation notation (contraction of indices) and
drop the explicit summation in the expression, i.e.
4
X
Λµν xν ≡ Λµν xν , (1.13)
ν=1

where a repeated pair of upper and lower indices implies a sum over that index such that

Λµν xν = Λµ0 x0 + Λµ1 x1 + Λµ1 x2 + Λµ1 x3 . (1.14)

As in the case of 3-vectors we can construct lengths that are invariant under Lorentz
transformations, in this case the invariant length is

A2 = (A0 )2 − (A1 )2 − (A2 )2 − (A3 )2 , (1.15)

for example

p2 = E 2 /c2 − |~p|2 = m2 c2 , pµ = (E/c, p~) . (1.16)

Put another way, the element in the 4D space

(ds)2 = (dt)2 − (dx1 )2 − (dx2 )2 − (dx3 )2 , (1.17)

is invariant under Lorentz transformations.


If we write this in terms of basis vectors

(Aµ êµ ) · (Aν êν ) = Aµ Aν êµ · êν , (1.18)

T. Blake 6 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

we can see that the basis vectors for µ = ν = 1, 2, 3 are not orthogonal. Instead êµ · êµ = −1.
At this point it is useful to introduce the (metric) tensor gµν and a new type of vector
xµ = (x0 , −~x). We call this type of vector a co-vector. The metric tensor in Minkowski
space is defined by

g00 = 1 , gii = −1 for i = 1, 2, 3 and gµν = 0 for µ 6= ν (1.19)

such that

(ds)2 = gµν dxµ dxν (1.20)

and similarly for g µν (which follows from gµν g µν = 1). The vectors and co-vectors are
related through the metric tensor,

xµ = gµν xν , (1.21)

which acts to lower and raise the indices. Anywhere we have a repeated upper and lower
index we have a dot product and get a scalar quantity, i.e. the dot product of A · B can
then be written as

Aµ Bµ = Aµ B µ = gµν Aµ B ν . (1.22)

Lorentz invariance implies

x0µ x0µ = gµν x0µ x0ν


(1.23)
= x0µ Λµν xν

∴ xν = x0µ Λµν , x0µ = [Λ−1 ]νµ xν (1.24)

i.e. the co-vectors transform with the inverse of the Lorentz transformation.
We can also write down the derivatives
   
µ ∂ ~ ∂ ~
∂ = , −∇ , ∂µ = ,∇ (1.25)
∂t ∂t

The derivative ∂µ is an important covector. Acting on a scalar field, φ,

∂φ ∂xν ∂φ
∂µ0 φ = =
∂x0µ ∂x0µ ∂xν (1.26)
= [Λ−1 ]νµ ∂ν φ

We want the 4-vector to be basis independent,

A = Aµ~eµ = A0ν ~eν0 . (1.27)

T. Blake 7 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

For this to hold,

Aµ~eµ = Λνσ Aσ Mνρ~eρ (1.28)


= Λνσ Mνρ Aσ~eρ

where M is some transformation of the basis vector. This in turn implies

Λνσ Mνρ = δσρ (1.29)

i.e. M is the inverse of the Lorentz transform, M = [Λ−1 ]. The basis vectors of Minkowski
space transform as

~eµ0 = [Λ−1 ]νµ~eν . (1.30)

We call Aµ a contravariant component and ~eµ a covariant basis vector. For the co-vectors,
the components are covariant and the basis vectors contravariant ,

A0µ = [Λ−1 ]νµ Aν (1.31)



~e = Λµν~e ν

In summary, to guarantee Lorentz covariance of our equations we need to write equations


that have the same type of object on both sides of the equation, e.g.
Aµ Bµ = k scalar ↔ scalar
Aµ Bµ C v = kC ν vector ↔ vector
Dµ Tµν = kCν vector ↔ vector

Aside: the light-cone


When we build quantum mechanical theories, it is important to remember that infor-
mation can not propagate faster than the speed of light. Events with space-time points
at xµ1 and xµ2 are only causal if (x1 − x2 )2 > 0. This is referred to as time-like separation.
Two events with (x1 − x2 )2 < 0 can not influence each other. The point x1 is not inside
the light-cone of x2 . This is referred to as space-like separation.

T. Blake 8 contact: thomas.blake@cern.ch


Lecture 2

Examples of Lorentz invariance –


Maxwell’s equations and the
Klein-Gordon equation

9
PP2 Relativistic Quantum Mechanics

2.1 Lorentz covariance of Maxwell’s equations


~ and a magnetic field B
Maxwell’s equations, written in terms of an electric field E ~ are

~ ·E
∇ ~ =ρ (Gauss’ law) ,
~
∇ ~ = − ∂B
~ ×E (Faraday-Lenz) ,
∂t (2.1)
~ ~
∇·B =0 (no magnetic charges) ,
~
∇ ~ = ~j + ∂ E
~ ×B (Ampere’s law) .
∂t
It turns out that these equations are Lorentz covariant, but this not immediately obvious
in this form. To Maxwell’s equations in a manifestly Lorentz covariant way, we start with
Ampere’s law and take its divergence
~
~ · (∇
∇ ~ × B)
~ =∇ ~ · ∂E
~ · ~j + ∇ (2.2)
∂t

~ · ~j + (∇ ~
~ · E)
0=∇ (2.3)
∂t
~ · ~j + ∂ρ .
=∇ (2.4)
∂t
This is the continuity equation, which we can express in terms of 4-vectors
 
∂ ~
∂µ = , ∇ , j µ = (ρ, ~j) , (2.5)
∂t

such that

∂µ j µ = 0 . (2.6)

The continuity equation implies that the charge


Z Z
3
Q= ρd x = j 0 d3 x (2.7)
V V

is locally conserved, i.e. the change in the charge in an arbitrarily small volume is only
due to the current flow out of the volume.
~ such that
At this point it is useful to introduce potentials, V and A,
~
~ =∇
B ~ ×A
~ , ~ = − ∂ A − ∇V
E ~ . (2.8)
∂t

T. Blake 10 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

~ ·B
By writing the magnetic and electric field in this way, the Faraday-Lenz law and ∇ ~ =0
~
are trivially satisfied. Re-writing Ampere’s law to include A and V
~ ×B
∇ ~ =∇
~ × (∇
~ × A)
~
!
∂ ~
∂A
= ~j + − ~
− ∇V (2.9)
∂t ∂t
~ ∇
= ∇( ~ · A)
~ −∇
~ 2A
~,

where we have used the triple product expansion

A × (B × C) = B(A · C) − C(A · B) . (2.10)

~ in Gauss’ law gives


Similarly, replacing E
∂ ~ ~ ~ 2V = ρ .
− (∇ · A) − ∇ (2.11)
∂t
Maxwell’s equations are also invariant under another type of transformation, Gauge
transformations. If we transform

~0 = A
~ − ∇χ , ∂χ
A V0 =V + , (2.12)
∂t
where χ is a scalar field, this does not change the magnetic or electric fields. We can
choose χ to simplify our expressions, by picking a solution that sets

∇ ~ + ∂V = 0 .
~ ·A (2.13)
∂t
~ then the Lorentz gauge is
This is the so-called Lorenz gauge. If we write, Aµ = (V, A),
manifestly Lorentz invariant,

∂µ A µ = 0 (2.14)

Back to
2
∂ ~ ~ ~ 2V = ρ ⇒ ∂ V − ∇
~ 2V = ρ
− (∇ · A) − ∇ (2.15)
∂t ∂t2
and
2~
~ ∇
∇( ~ · A)
~ −∇ ~ = ~j − ∂ A − ∇
~ 2A ~ ∂V (2.16)
∂t 2 ∂t
which gives
~
∂ 2A ~ 2A
−∇ ~ = ~j . (2.17)
∂t 2

T. Blake 11 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

This can be written in the compact form

∂ µ ∂µ A ν = j ν (2.18)

where
∂2 ~2 .
∂µ ∂ µ = 2
−∇ (2.19)
∂t
This expression is also manifestly Lorentz covariant. You will sometimes see this written
with the replacement

 ≡ ∂µ ∂ µ so Aµ = j µ . (2.20)

In summary, we can write Maxwells equations in a compact, Lorentz covariant form as

∂µ j µ = 0 , ∂ µ ∂µ A ν = j ν (2.21)

with

j µ = (ρ, ~j) , ~ and ∂µ Aµ = 0 .


Aµ = (V, A) (2.22)

T. Blake 12 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

2.2 Quantum mechanics


So far we have focussed on special relativity. Now we want to see how this can be combined
with quantum mechanics (QM). The fundamental building blocks we need from quantum
mechanics are:

• Observables can be continuous or discrete.

• Variables are associated to operators, i.e. we can promote

E, x, p → Ê, x̂, p̂ . (2.23)

• Systems can be described by state vector |ψi in Hilbert space.

• The operators act on the state vectors

Â|ψi = a|ψi (2.24)

where a is an eigenvalue and |ψi and eigenvector.

• The eigenvectors can be made orthogonal such that

hψm |ψn i = δmn , h~x1 |~x2 i = δ 3 (~x1 − ~x2 ) (2.25)

• The eigenvectors form a complete set that spans the space such that
X
|ψm ihψm | = 1 (2.26)
m
Z
d3~x|~xih~x| = 1 (2.27)

• From the eigenvectors, we can define position-space wave-functions

ψ(~x) = h~x|ψi (2.28)

which project a state ψ onto a position ~x.

• In the Born interpretation, ψ(~x) is interpreted in terms of position probability density


ρ(~x) = ψ ∗ ψ
Z
ψ ∗ (~x)ψ(~x)d3~x = 1 (2.29)

• In the Schrödinger picture the state vectors are time-dependent,

|ψi → |ψ(t)i and ψ(~x, t) = h~x|ψ(t)i (2.30)

T. Blake 13 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

2.3 The Scrödinger equation


In classical mechanics
|~p|2
E= + V (~x, t) (2.31)
2m
Promoting E and p to to operators gives

E → Ê = i~ , (2.32)
∂t
p~ → p̂ = −i~∇~ ,

Dropping the ~, these can now be written in a neater 4-vector notation as

pµ = i∂ µ . (2.33)

The operators Ê and p̂ operate on a wave-function ψ(~x, t) and this yields the Schrödinger
equation
∂ ~2 ~ 2
i~ ψ(~x, t) = − ∇ ψ(~x, t) + V (~x, t)ψ(~x, t) , (2.34)
∂t 2m
where V (~x, t) is an arbitrary potential.
We can also define a particle density by

ρ = ψ∗ψ (2.35)

Differentiating ρ by parts yields


∂ρ ∂ψ ∂ψ ∗
= ψ∗ + ψ. (2.36)
∂t ∂t ∂t
Using the complex conjugate of the Schrödinger equation
∂ ∗ ~2 ~ 2 ∗
−i~ ψ (~x, t) = − ∇ ψ (~x, t) + V ∗ (~x, t)ψ ∗ (~x, t) , (2.37)
∂t 2m
and assuming V = V ∗ gives
∂ψ i ∗~ 2 ∂ ∗ψ i ~2 ∗
ψ∗ − ψ ∇ ψ = −ψ − ψ∇ ψ , (2.38)
∂t 2m ∂t 2m
which can be written (you will do something similar as an exercise) as
∂ρ ~ ~
+ ∇ · j = ∂µ j µ = 0 . (2.39)
∂t
Here, ρ is the probability density and ~j is a probability current and j µ = (ρ, ~j). This is
the continuity equation for a conserved quantity (conserved probability in this case).

T. Blake 14 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

2.4 The Klein-Gordon equation


The Schrödinger equation was derived from the non-relativistic energy-momentum relation.
It’s non linear in E and p~, which does not allow us to use a Lorentz covariant notation.
We know that there is a more appropriate relativistic equation. What happens if we take
E 2 = |~p|2 + m2 (2.40)
and promote E and p~ to operators? This gives
 2 
∂ ~ 2 2
⇒ − ∇ + m φ(~x, t) = 0 (2.41)
∂t2
which can be written in a compact notation as
(∂ µ ∂µ + m2 )φ(x) = 0 . (2.42)
This is the Klein-Gordon equation. The Klein-Gordon equation has plane wave solutions
of the form
~
φ(x) = N e−i(ωt−k·~x) = N e−ip·x (2.43)
Plugging φ(xµ ) back into the equation yields
p
E = ± |~p|2 + m2 (2.44)
This is the equation for a free spin-0 particle. There are both positive and negative energy
solutions (which should not come as a big surprise because we started with an expression
for E 2 ).
Multiplying the Klein-Gordon equation by −iφ∗ and its complex conjugate by −iφ,
gives
∂ 2φ ∗~ 2 ∂ 2 φ∗ ~ 2 φ∗ − im2 φφ∗ = 0
−iφ∗ + iφ ∇ φ − im 2 ∗
φ φ = −iφ + iφ∇ (2.45)
∂t2 ∂t2
subtracting the two expressions
∂φ∗
 
∂ ∗ ∂φ

~ · φ∗ ∇φ~ − φ∇φ

~ ∗ =0
i φ −φ − i∇ (2.46)
∂t ∂t ∂t
this is the continuity equation for the Klein-Gordon equation,
∂µ j µ = 0 (2.47)
with
j µ = i(φ∗ ∂ µ φ − φ∂ µ φ∗ ) . (2.48)
Plugging in the plane-wave solutions of the Klein-Gordon equation gives
j µ = 2|N |2 pµ , pµ = (E, p~) (2.49)
We now have a second problem, the negative energy solutions also have a negative
probability density since j 0 = 2|N |2 E. Unfortunately, the problem can’t be simply ignored
because in QM we need to work with a complete set of states.

T. Blake 15 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

2.5 The Feynman-Stückleberg interpretation


A prescription for handling negative energy states was proposed by Stückleberg (1941)
and Feynman (1946):

Negative energy solutions describe a particle propagating backwards in time, or,


a positive energy anti-particle propagating forwards in time.

Suppose the Klein-Gordon equation describes a free spin-0 particle of charge +e, e.g. a
π + . Then the electromagnetic-current associated with the charged particle is

j µ (π + ) = 2e|N |2 (E, p~) (2.50)

A π − with the same energy an momentum has a current

j µ (π − ) = −2e|N |2 (E, p~) (2.51)

which is the same current as the π + under E → −E and p~ → −~p. A negative-energy


particle solution going backwards in time describes a positive energy anti-particle solution
going forward in time. The reason this works is because e−i(−E)(−t) = e−iEt . Whilst the
Klein-Gordon wave-functions are fundamentally single particle solutions, the Feynman-
Stückleberg interpretation allows us to handle many particle states.

T. Blake 16 contact: thomas.blake@cern.ch


Lecture 3

Perturbation theory and


particle scattering

17
PP2 Relativistic Quantum Mechanics

3.1 Time dependent perturbation theory


In High Energy Physics we deal with situations that can be described by
1. free incoming particles of known (E, p~ ) (beam);
2. an interaction;
3. and free outgoing particles of known (E, p~ ) (measured in a detector).
The interaction acts over a short time T and can be described by a potential V (t, ~x), where
V (t, ~x) = 0 outside of T . The interactions are also weak, such that the probability of an
interaction occurring P(interaction)  1. The smallness of the interaction allows us to
make perturbative expansions.

Figure 3.1: Illustration of a typical scattering problem in High Energy Physics.

We will use a non-relativistic framework to derive the tools we need to do scattering


calculations. This might sound questionable, but a full treatment gives the same result. If
you want to understand the reasons behind this, it is explained in Relativistic Quantum
Fields by Bjorken and Drell.
We will start with the Schrödinger equation for a free particle. This is given by
i∂φ
= H0 (~x)φ (3.1)
∂t
where the Hamiltonian, H0 , is constant in time, i.e.

H0 (~x)φm (~x) = Em φm (~x) , (3.2)

but the states are time dependent,

φm (t, ~x) = φm (~x)e−iEm t . (3.3)

T. Blake 18 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

The states are normalised such that


Z
φ∗m (~x)φn (~x)d3~x = δmn . (3.4)

The wave-functions φ form a complete basis so the solutions to this equation can be
expressed as
X
ψ= am (t)φ(~x)e−iEm t (3.5)
m

The objective is then to solve


∂ψ
[H0 + V (~x, t)] ψ = i , (3.6)
∂t
where we have introduced a time-dependent potential V (~x, t). The normalisation of the
wave-function ψ is given by
Z X
ψ ∗ ψ d4 x = 1 implies |am |2 = 1 , (3.7)
m

where |am |2 is the probability to find the particle in state φm . This is time-independent
for a free system.
After we have introduced the potential
∂ψ
[H0 + V (~x, t)] ψ = i (3.8)
∂t
∂ X
=i am (t)φ(~x)e−iEm t ,
∂t m

the am (t) are time-dependent and we can get transitions between states. This can be
expanded as
X
[H0 + V (~x, t)] am (t)φ(~x)e−iEm t =
m
X  ∂am (t)  (3.9)
i − iEm am (t) φ(~x)e−iEm t
m
∂t

giving
" #
X ∂am (t) X
i φ(~x)e−iEm t = V am (t)φ(~x)e−iEm t (3.10)
m
∂t m

where we have used

H0 φm (~x) = Em φm (~x) . (3.11)

T. Blake 19 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

We can simplify this expression by multiplying through by φ∗f (~x) and integrating over the
phasespace, i.e.
Z X
∂am (t) ∗
i φf (~x)e+iEf t φm (~x)e−iEm t d3~x
m
∂t
Z X (3.12)
∗ +iEf t −iEm t 3
= am (t)φf (~x)e V (t, ~x)φ(~x)e d ~x
m

and simplify further using


Z
φ∗f (~x)φm (~x)d3~x = δf m . (3.13)

This gives us a series of coupled equations


Z
∂af (t) X
= −i am (t)ei(Ef −Em )t φ∗f (~x)V (t, ~x)φm (~x)d3~x (3.14)
∂t m

from which we can define


Z
Vf m = φ∗f (~x)V (t, ~x)φm (~x)d3~x . (3.15)

It is not possible to solve the system exactly, but we can exploit the weakness of the
potential and choose to pull out a small scalar coupling constant κ, i.e. to replace V (t, ~x)
by κV (t, ~x) and Vf m by κVf m . We can also express the am (t) as a power series in terms of
this small coupling constant as
X
am (t) = (κ)j ajm (t) , (3.16)
j

where
da0m
=0. (3.17)
dt
After comparing powers of κ, we have

daj+1
f (t)
X
= −i Vf m ei(Ef −Em )t ajm (t) (3.18)
dt m

At zeroth order
da0f (t)
= 0 ⇒ af = δf i (3.19)
dt

T. Blake 20 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

i.e. there is no interaction. At first order


da1f (t) X
= −i Vf m ei(Ef −Em )t a0f (3.20)
dt m
X
= −i Vf m ei(Ef −Em )t δmi , (3.21)
m

where we have substituted in the zeroth order result. This gives


Zt
0
a1f (t) = −i Vf i ei(Ef −Ei )t dt0 . (3.22)
−∞

At second order
da2f (t) t
Z
0
X
= −i dt0 Vf m ei(Ef −Em )t a1m (t0 ) (3.23)
dt m −∞

Zt Z t0
i(Ef −Em )t0 0 00
a2f (t) = (−i)2
Vf m e dt Vmi ei(Em −Ei )t dt00 . (3.24)
−∞
−∞

where we have now substituted in the first order result. This process can then be repeated
for higher-and-higher orders in κ. It is also important to note that this process is time-
ordered.
If the potential is time-independent, i.e. V = V (~x), then
X 1
a2f (t) = −2πiδ(Ef − Ei ) Vf m Vmi , (3.25)
m6=i
Ei − Em + iε

which is integrated by introducing a small complex parameter Em − Ei → Em − Ei − iε


(where  → 0). The zeroth, first and second order processes are shown schematically in
Fig. 3.2. The first order contribution gives us an interaction with f 6= i.

Figure 3.2: Zeroth, first and second-order contributions to the i → f transition.

T. Blake 21 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

As a simplification we can take the system to be in state i at t = −T /2

ai (−T /2) = 1 (3.26)


an (−T /2) = 0 for n 6= i

which gives
Z
daf
= −i φ∗f (~x)V φi (~x)ei(Ef −Ei )t d3~x (3.27)
dt
and at time T /2 after the interaction has ceased,

ZT /2
Tf i ≡ af (T /2) = −i φ∗f (~x)V φi (~x)ei(Ef −Ei )t d3~xdt (3.28)
−T /2

or more generally in a covariant form


Z
Tf i = af (t) = −i φ∗f (x)V (x)φi (x)d4 x (3.29)

We would like to interpret |Tf i |2 as the probability that a particle is scattered from state i
into state f . For the time-independent potential
Z T /2
Tf i = −iVf i ei(Ef −Ei )t dt (3.30)
−T /2

such that
sin2 ( T2 (Ef − Ei ))
|Tf i |2 = 4|Vf i |2 (3.31)
(Ef − Ei )2
This function is tightly peaked around Ef = Ei . If we take T → ∞

Tf i = 2πiVf i δ(Ef − Ei ) (3.32)

the δ-function expresses energy conservation in the transition. Typically we want to know
the transition rate from a known initial state to a group of final states (e.g. a particles
decay width),
X Z Z
2
|Ttot,f i | = |Tf i | = |Tf i | dN (Ef ) = |Tf i |2 ρ(Ef )dEf
2 2
(3.33)
f

Assuming ρ(Ef ) and Vf i are constant over a narrow integration window,

sin2 ((Ef − Ei ) T2 )
Z
2 2
|Ttot,f i | = 4ρ(Ef )|Vf i | dEf (3.34)
(Ef − Ei )2

T. Blake 22 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Figure 3.3: |Tf i |2 for fixed T as a function of Ef − Ei . It is sharply peaked at Ef ≈ Ei .

lim |Ttot,f i |2 = 2πT |Vf i |2 ρ(Ef ) (3.35)


T →∞

T appears on the right-hand side so this is not a transition rate. We can instead define a
transition rate per unit time as

|Ttot,f i |2
W = lim = 2π|Vf i |2 ρ(Ef ) (3.36)
T →∞ T
This is Femi’s golden rule. We made a lot of approximations to get here, but as long as the
interaction is weak this rule holds. In the QFT course, you will re-visit time-dependent
perturbation theory using operators and the Dirac interaction picture.

T. Blake 23 contact: thomas.blake@cern.ch


Lecture 4

Coulomb scattering of
charged spin-0 particles

24
PP2 Relativistic Quantum Mechanics

4.1 Motion of a charged particle in an EM field


We have already seen that free spin-0 particles can be described by the Klein-Gordon
equation

(∂ µ ∂µ + m2 )φ = 0 . (4.1)

In classical EM, the motion of a particle with charge (−e) in a field Aµ can be obtained
by substituting

pµ → pµ + eAµ (4.2)

The QM analogue is then

i∂ µ → i∂ µ + eAµ (4.3)

This is Quantum-Electro-Dynamics (QED). Plugging this into the KG equation gives


 µ
(∂ − ieAµ )(∂µ − ieAµ ) + m2 ) φ = 0

 µ (4.4)
∂ ∂µ − e2 Aµ Aµ − ie(∂ µ Aµ + Aµ ∂µ ) + m2 φ = 0


where ie∂ µ Aµ acts on φ, i.e. ie∂ µ (Aµ φ). This can be written in the form of an equation of
motion that depends on a potential

(∂ µ ∂µ + m2 )φ = −V φ , (4.5)

where

V = −ie(∂ µ Aµ + Aµ ∂µ ) − e2 Aµ Aµ . (4.6)

This is the interaction potential for the charged particle in the field. Since e  1,

e2 1
αEM = ∼ , (4.7)
4π 137
we can simplify the expression by taking only the first order terms

V ≈ −ie(∂ µ Aµ + Aµ ∂µ ) (4.8)

We now want to compute the transition current


Z
Tf i = −i φ∗f (x)V (x)φi (x)d4 x (4.9)
Z
Tf i = i2
φ∗f (x)e(Aµ ∂µ + ∂µ Aµ )φi (x)d4 x

T. Blake 25 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Figure 4.1: A spineless particle interacting with a field Aµ .

This can be integrated by parts


Z+∞ +∞ Z+∞
∗ ∗
µ 4
φf (x)∂ (Aµ φi (x))d x = φf Aµ φi − (∂ µ φ∗f (x)Aµ φi (x)d4 x (4.10)

−∞
−∞ −∞

Assuming Aµ → 0 as xµ → ∞
+∞
φ∗f Aµ φi →0 (4.11)

−∞

Giving
Z
Tf i = −i (−ie)(φ∗f ∂µ φi − φi ∂µ φ∗f )Aµ d4 x (4.12)

(4.13)

This should look familiar, it is the Klein-Gordon current, only having different states for
φi and φf . From this we can define a transition current

jµf i = −ie(φ∗f ∂µ φi − φi ∂µ φ∗f ) (4.14)

and write
Z
Tf i = −i jµf i Aµ d4 x (4.15)

(4.16)

Now we can look at the initial and final states. These are both free spin-0 particles,
which obey the Klein-Gordon equation with solutions

φ = N e−ip·x (4.17)

Using

∂µ φ = −ipµ φ (4.18)

the current takes the form

jµf i = −e(pi + pf )µ φ∗f φi (4.19)

T. Blake 26 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

The final ingredient we need is the field Aµ . For coulomb scattering the source of this field
is the other particle. We already know that there will be a transition current associated
(2)
with this particle, jµ , and can use Maxwell’s equations to find an expression for the field,
µ
∂ µ ∂µ Aµ = j(2) = −ie(φ∗D ∂ µ φB − φB ∂ µ φ∗D ) (4.20)
= −e(pB + pD )µ NB ND ei(pD −pB )·x

We often write pD − pB = q.

Figure 4.2: Two particles scattering off each other. The particle four-momentum are indicated
by pA , pB , pC and pD . For coulomb scattering the internal line corresponds to the exchange of a
photon.

By realising that

∂ µ ∂µ e−iq·x = −q 2 e−iq·x (4.21)


µ
we can express Aµ in terms of j(2) as
µ
µ
j(2)
A =− . (4.22)
q2
Finally, putting the ingredients together
Z  
(1) 1 µ
Tf i = −i jµ − 2 j(2) d4 x (4.23)
q
and plugging in the the plane wave solutions to the Klein-Gordon equation

jµ(1) = −eNA NC (pA + pC )µ ei(pC −pA )·x (4.24)


µ i(pD −pB )·x
j(2) = −eNB ND (pB + pD )µ e
(4.25)

T. Blake 27 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

gives
1
Tf i = −ie2 NA NB NC ND (pA + pC )µ 2 (pB + pD )µ
q
Z (4.26)
× ei(pC −pA +pD −pB )·x d4 x

The integral results in the δ-function


Z
ei(pC −pA +pD −pB )·x d4 x = (2π)4 δ 4 (pC − pA + pD − pB ) , (4.27)

which ensures 4-momentum conservation. We can also pull out an Lorentz invariant
amplitude
ig µν
 
−iM = ie(pA + pC )µ − 2 ie(pB + pD )ν (4.28)
q
It is interesting to note that in this process q 2 =
6 0, which implies that mγ 6= 0. The photon
is a “virtual” or “off-mass-shell” particle.

4.2 Feynman rules for spin-0 particles


The building blocks we need to perform Coulomb scattering calculations are:

Figure 4.3: Feynman rules for charged spin-0 particles

External photons are accompanied by a polarisation vector ε∗µ (p, λ), where p is the
photon momentum and λ is the spin projection of the photon along p~. We won’t derive the

T. Blake 28 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

origin of this factor in these notes. There are also two different internal lines depending
on whether a photon or a spin-0 particle are exchanged. We refer to these internal lines as
propagators.

T. Blake 29 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

4.3 Propagators and time-ordering


If we look again at the diagram in Fig 4.2 we actually have a diagram with two vertices,
i.e.
−iM ∝ e2 . (4.29)
This process should correspond to a second order term in the perturbative expansion.
Recalling, Eq. 3.25, this should be related to
X 1
−2πiδ(Ef − Ei ) Vf m Vmi . (4.30)
Ei − Em + iε
m6=i

The term
1
(4.31)
Ei − Em
in the second order expression looks a lot like a propagator. To make the connection to
the propagator its important to realise that when we draw a Feynman diagram we really
mean a sum over different possible time-ordered diagrams. When we do the calculation
we integrate over space-time and enforce energy conservation for the external particles.

Figure 4.4: Feynman diagrams are a sum over the different possible time-ordered diagrams.

For the first time ordering in Fig. 4.4, the energy of the intermediate state is
p
Em = EC + EB + |~pA − p~C |2 + m2 , (4.32)
p
where the exchanged particle has energy EP = |~pA − p~C |2 + m2 and mass m. The energy
of the initial state is
Ei = EA + EB . (4.33)

T. Blake 30 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

This gives
p
Ei − Em = (EA + EB ) − EC − EB − |~pA − p~C |2 + m2 (4.34)
p
= (EA − EC ) − |~pA − p~C |2 + m2 .

The same exercise for the second time ordering in Fig. 4.4 gives
p
Em = EA + ED + |~pB − p~D |2 + m2 (4.35)

and
p
Ei − Em = −(EA − EC ) − |~pA − p~C |2 + m2 . (4.36)

Putting the two parts together


p
1 1 2 |~pA − p~C |2 + m2 2 EP
+ = 2 2
= 2 (4.37)
(Ei − Em )(1) (Ei − Em )(2) (pA − pC ) − m q − m2

The transition amplitude, Tf i , is then proportional to the propagator we expected. You


can also see how this relates to the propagator for an internal spin-0 particle.

T. Blake 31 contact: thomas.blake@cern.ch


Lecture 5

Observable quantities

32
PP2 Relativistic Quantum Mechanics

5.1 Box normalisation


From the Klein-Gordon equation, the plane wave solutions for a free particle are

φ = N e−ip·x , (5.1)

which correspond to a current

j µ = 2|N |2 pµ (5.2)

such that the density is

ρ = j 0 = 2E|N |2 . (5.3)

The proportionality of ρ to E is needed to compensate for the Lorentz contraction of the


volume element d3~x, leaving the number of particles ρd3~x unchanged. We usually choose
a normalisation of 2E particles in the arbitrary normalisation volume, V ,
2E 1
ρ= ⇒ N=√ . (5.4)
V V
We will see later that V drops out in our calculations and you will often see this written
as ρ = 2E. The density of states is given by

dn = ρ(E)dE . (5.5)

Figure 5.1: Box normalistion of states. States exist in box with sidelength L.

T. Blake 33 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

If the states exist in a box of side L, we know from the infinite square well problem in
Quantum Mechanics that in one dimension (x),

px = nx , (5.6)
L
where nx is an integer. This gives

dpx = dnx . (5.7)
L
The total number of states in a cube of side L with momentum between p~ and p~ + d3 p~ is
then

dn = dnx dny dnz (5.8)


 3
L
= dpx dpy dpz

 3
L V
= d3 p~ = d3 p~
2π (2π)3

We have 2E particles in the volume V , so the number of states per particle is


V
dn = d3 p~ . (5.9)
(2π)3 2E

T. Blake 34 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

5.2 Computing observable quantities from


invariant amplitudes
The task is now to turn the Lorentz covariant amplitude M into a measurable quantity.
In particle physics we typically want to compute:

1. Cross-sections, σ, for a process (e.g. A + B → C + D)

2. Decay rates, Γ, for a particle (e.g. A → B + C)

To compute σ or Γ, we start with Tf i written in terms of M


Y Y
Tf i = −iMδ 4 (pf − pi ) Nf Ni (5.10)
f i

Here, pf and pi are the final and initial state 4-momentum summing over all of the particles.
The Nf and Ni are the normalisations associated to each of the initial and final state
particles. The δ-function ensures energy and momentum conservation in the process.

Figure 5.2: Illustration of a typical scattering problem in High Energy Physics, occuring in a
local volume V over time T .

We can turn this amplitude into a probability for a transition from i → f by taking its
square,
2
Y Y
pf i = |Tf i |2 = − iMδ 4 (pf − pi ) Nf Ni . (5.11)


f i

T. Blake 35 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

This includes the square of a δ-function. To deal with this can replace the δ-function with
the identity
Z
(2π) δ (pf − pi ) = ei(pf −pi )·x d4 x ,
4 4
(5.12)

such that
Z
4 4 2 4 4
|(2π) δ (pf − pi )| ≈ (2π) δ (pf − pi ) ei(pf −pi )·x d4 x (5.13)
T,V

where the integral is over the volume of the interaction region, V , and the extent of the
interaction in time, T . The first δ-function forces pf = pi , so this is simply
|(2π)4 δ 4 (pf − pi )|2 ≈ V T (2π)4 δ 4 (pf − pi ) (5.14)
Putting the ingredients back into |Tf i |2 ,
Y Y
|Tf i |2 = |M|2 V T (2π)4 δ 4 (pf − pi ) |Nf |2 |Ni |2
f i (5.15)
1 1
= |M|2 V T (2π)4 δ 4 (pf − pi ) nf ni
V V
where nf and ni are the number of initial and final state particles. For a process involving
A + B → C + D, we would have
|Tf i |2 = |M|2 V T (2π)4 δ 4 (pC + pD − pA − pB )|NA |2 |NB |2 |NC |2 |ND |2 (5.16)
1
= |M|2 V T (2π)4 δ 4 (pC + pD − pA − pB ) 4
V
The transition rate per unit volume can be computed from |Tf i |2 as
|Tf i |2 1 1
Wf i = = |M|2 (2π)4 δ 4 (pf − pi ) nf ni . (5.17)
VT V V
The transition rate per unit volume into a small fixed final state phase-space is
dW = Wf i dn (5.18)
1 1 Y d3 p~
= |M|2 ni fi (2π)4 δ 4 (pf − pi ) V
V V f
2Ef (2π)3
1 Y d3 p~
= |M|2 (2π)4 4
δ (p f − p i )
V ni f
2Ef (2π)3
(5.19)
We call
Y d3 p~
dQ = (2π)4 δ 4 (pf − pi ) (5.20)
f
2Ef (2π)3

the Lorentz invariant phase-space.

T. Blake 36 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

5.3 Decay rates


Lets now consider a decay A → B + C and compute
decays per unit time
decay rate = (5.21)
decaying particles
from the number of transitions per unit volume per unit time, dW . The number of particles
of type A per unit volume is 2EA /V , yielding the differential decay rate
V
dΓ = dW (5.22)
2EA
V 1 d3 p~B d3 p~C
= |M|2 (2π)4 δ 4 (pA − pB − pC ) .
2EA V 2EB (2π)3 2EC (2π)3
The size of the arbitrary normalisation volume, V , drops out. The differential decay rate,
dΓ is covariant so we can pick a frame to make this calculation easier. In the rest frame of
A,
pA = (mA , 0) , pB = (EB , p~ ) , pC = (EC , −~p ) . (5.23)
The decay rate is then
d3 p~B d3 p~C
Z
1 2 4
Γ= |M| δ (p A − p B − p C ) . (5.24)
8(2π)2 mA EB EC
We can integrate over one of the final state momenta using the δ-function to fix p~B = −~pC .
This gives
d3 p~
Z
1 2
Γ= |M| δ(mA − EB − EC ) . (5.25)
8(2π)2 mA EB EC
To put this into a more usable form we can use polar coordinates
d3 p~ = |~p |2 d|~p |dΩ . (5.26)
We can also use the relativistic energy momentum relationship
E 2 = |~p|2 + m2 (5.27)
to write
|~p|d|~p| |~pB |d|~pB | |~pC |d|~pC |
dE = and d(EB + EC ) = + (5.28)
E EB EC
but since |~pB | = |~pC |,
|~p |d|~p| |~p |d|~p|
d(EB + EC ) = + (5.29)
EB EC
(5.30)

T. Blake 37 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

rearranging gives
EB EC
|~p|d|~p| = d(EB + EC ) . (5.31)
EB + EC
We can use this to write
d3 p~ |~p|2 d|~p|dΩ |~p|d(EB + EC )dΩ
= = , (5.32)
EB EC EB EC EB + EC
and get

|~p|d(EB + EC )dΩ
Z
1
Γ= |M|2 δ(mA − EB − EC ) (5.33)
8(2π)2 mA EB + EC

which can again be integrated by exploiting the δ-function to fix mA = EB + EC . Finally


we arrive at
Z
1
Γ= |M|2 |~p|dΩ (5.34)
32π 2 m2A

5.4 Cross sections


We can also compute the cross-section for a scattering process A + B → C + D. Lets
start by considering a beam of particles A incident on a target of type B. The number of
particles of type A passing through unit area per unit time is
2EA
|~vA | , (5.35)
V
where ~vA is the velocity of the particles in the beam of A,
p~A
~vA = . (5.36)
EA
The number of particles per unit volume in the target is
2EB
. (5.37)
V
The cross-section for the process is then given by
number of final states
σ(A + B → C + D) = Wf i × , (5.38)
initial flux
where we have

2 4 4 |M|2
Wf i = |Tf i | = (2π) δ (pC + pD − pA − pD ) 4 (5.39)
V

T. Blake 38 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

The initial flux of particles in the lab-frame is


2EA 2EB
|~vA | , (5.40)
V V
and the number of final states between p~C and p~C + d3 p~C and between p~D and p~D + d3 p~D
is
V d3 p~C V d3 p~D
. (5.41)
(2π)3 2EC (2π)3 2ED

This gives

1 1 4 d3 p~C d3 p~D
dσ = |M|2 δ (p C + p D − p A − p D ) , (5.42)
4EA EB |~vA | (2π)2 2EC 2ED

where once again the arbitrary normalisation volume cancels. It’s useful to write this in
terms of the Lorentz invariant phase-space

|M|2
dσ = dQ , (5.43)
F
where F is the incident flux

F = |~vA |2EA 2EB . (5.44)

More generally, for a collinear collision between A and B,

F = |~vA − ~vB |2EA 2EB . (5.45)

From which can can derive,

dσ 1 |~pC |
= 2 2
|M|2 . (5.46)
dΩ 64π (pA + pB ) |~pA |

5.5 Mandelstam variables


At this point it is useful to introduce the kinematic variables

s = (pA + pB )2 = (pC + pD )2
t = (pA − pC )2 = (pB − pD )2 (5.47)
u = (pA − pD )2 = (pB − pC )2

These are called Mandelstam variables. They will appear in scattering calculations and
are Lorentz covariant, they can be used to describe a process in a frame independent way.

T. Blake 39 contact: thomas.blake@cern.ch


Lecture 6

Calculating cross-sections
for spin-0 particle scattering

40
PP2 Relativistic Quantum Mechanics

6.1 Spinless scattering: e−µ− → e−µ−


The procedure to compute a scattering cross-section is the same for all processes:
1. At a given order, draw all possible Feynman diagrams for the process.

2. Using the Feynman rules, label the elements in each diagram with the appropriate
factor.

3. Form an invariant amplitude −iM by multiplying all of the factors in the proper
order. This is the order which follows the lines in the diagram against the direction
of the particle flow.

4. Compute the transition probability from |M|2 .


The transition probability can then be used to calculate cross-sections or decay rates using
the formulae we derived earlier.
The first concrete example we’ll work through is spin-0 e− µ− → e− µ− scattering. This
has a single Feynamn diagram at the lowest order, see Fig. 6.1. In reality, we know that
electrons and muons are both spin- 12 particles and we’ll do the full calculation later in this
course.

Figure 6.1: Feynman diagram for electron-muon scattering.

From the diagram we can write the invariant transition amplitude as


−ig µν
 
−iMe− µ− = [1][ie(pA + pC )µ ][1] [1][ie(pB + pD )ν ][1] (6.1)
q2
1
= ie2 (pA + pC )µ (pB + pD )µ ,
(pA − pC )2

T. Blake 41 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

where we have used q = pA − pC . If there were more diagrams, we would sum over them
in the amplitude. The transition probability
2
(pA + pC )µ (pB + pD )µ

2 4
|Me− µ− | = e (6.2)
(pA − pC )2
 2
4 pA · pB + pA · pD + p C · pB + pC · pD
=e
t

We can now write this expression in terms of the Mandelstam variables

s = (pA + pB )2 = (pC + pD )2
= p2A + p2B + 2pA · pB = m2A + m2B + 2pA · pB
= p2C + p2B + 2pC · pD = m2C + m2D + 2pC · pD
(6.3)
2 2
u = (pA − pD ) = (pB − pC )
= p2A + p2D − 2pA · pD
= p2B + p2C − 2pB · pC ,
as

pA · pB + pA · pD + pC · pB + pC · pD = (6.4)
1
2
(s − m2A − m2B ) + 12 (s − m2C − m2D ) − 12 (u − m2A − m2D ) − 12 (u − m2B − m2C )

such that
 2
2 4 s−u
|Me− µ− | = e .
t

This finally brings us to the expression for the differential cross-section


2
e4 |~pC | s − u


= . (6.5)
dΩ 64π 2 s |~pA | t

In the centre of mass system |~pA | = |~pC |. Further, in the very high energy limit me  Ee
and mµ  Eµ and the differential cross-section becomes
2
e4

dσ 3 + cos θ
= , (6.6)
dΩ 64π 2 s 1 − cos θ
CM

where, θ is the angle between p~A and p~C . The cross-section diverges as θ → 0.

T. Blake 42 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Figure 6.2: Centre of mass system for e− µ− → e− µ− scattering.

6.2 Spinless scattering: e−e− → e−e−


Now lets consider a slightly more complex situation where we have electron-electron
scattering, e− e− → e− e− . The complication here is that we have identical particles in the
initial and final state. We therefore cannot distinguish C from D (and A from B). The
resulting amplitude should be symmetric under the exchanges A ↔ B and C ↔ D. The
invariant amplitude is the sum of the two diagrams in Fig. 6.3. The first diagram is the
same one we had for e− µ− → e− µ− scattering. The second diagram interchanges C and
D.

Figure 6.3: Diagrams for e− e− → e− e− scattering. Note, in the second diagram the two lines do
not cross.

The invariant amplitude corresponding to these diagrams is

(pA + pC )µ (pB + pD )µ 2 (pA + pD )µ (pB + pC )


µ
−iMe− e− = ie2 + ie . (6.7)
(pA − pC )2 (pA − pD )2

It’s easy to see that this is invariant under C → D. It’s also invariant under B → D. This

T. Blake 43 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

can be written using the Mandelstam variables as


 
2 s−u s−t
−iMe− e− = ie + (6.8)
t u

6.3 Infrared divergences


The invariant amplitude for e− e− → e− e− scattering diverges as t → 0 and u → 0. This
corresponds to a divergence at scattering angles close to zero and π. The Mandelstam
variables t and u in this case correspond to the 4-momentum transferred by the virtual
photon and t(u) → 0 implies q 2 → 0 such that the photon approaches the on mass-shell
condition. To solve the problem of divergences in the theory, it is important to both
include higher order diagrams and to renormalise the theory. You will come across this in
later courses.

Figure 6.4: Differential cross-section for e− e− → e− e− scattering as a function of the scattering


angle, θ.

T. Blake 44 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

6.4 Spinless scattering: e+e− → e+e−


Up-to now we have just looked at particle-particle scattering. Now let’s try to compute
e+ e− → e+ e− scattering. In this case the particles are distinguishable in the initial
and final state, but we again have two diagrams. The second diagram corresponds to
particle-antiparticle annihilation. The two diagrams are shown in Fig. 6.5.

Figure 6.5: Feynman diagrams for spinless electron-positron scattering.

In order to write down the invariant amplitude we need Feynman rules involving an-
tiparticles. If we recall the Feynman-Stückelberg picture then a positive energy antiparticle
travelling forwards in time is equivalent to a negative energy particle travelling backwards
in time.
We can use the Feynamn-Stückelberg picture for example to write the vertex term
for the lower vertex in the first diagram as ie(−pB − pD )ν . For the first diagram, the
4-momentum of the exchanged photon as q = pA − pC . For the second diagram, the
4-momentum of the photon is q = pA + pB . From the diagrams, the invariant amplitude is
then
µ µ
2 (pA + pC )µ (−pB − pD ) 2 (pA − pB )µ (pC − pD )
−iMe e = ie
+ − + ie (6.9)
(pA − pC )2 (pA + pB )2
Comparing this to the expression we had earlier we find that
Me− e− (pA , pB , pC , pD ) = Me+ e− (pA , −pD , pC , −pB ) (6.10)
i.e. it is the same expression just exchanging pB ↔ −pD . We can exploit this to short cut
to the result
s = (pA + pB )2 → u
t = (pA − pC )2 → t (6.11)
u = (pA − pD )2 → s

T. Blake 45 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

such that
 
2 u−s u−t
−iMe+ e− = ie + (6.12)
t s

We refer to the diagram that contributes (u − s)/t to the amplitude as a t-channel process
and the diagram that contributes (u − t)/s as an s-channel process.

T. Blake 46 contact: thomas.blake@cern.ch


Lecture 7

The Dirac equation

47
PP2 Relativistic Quantum Mechanics

7.1 The Dirac equation


The Klein-Gordon equation for a free, spin-0, particle with wave-function φ is given by
∂ 2φ ~ 2
2
− ∇ φ + m2 φ = 0 . (7.1)
∂t
When deriving this expression, we came across two problems:
1. The equation has solutions with E < 0,
2. The E < 0 solutions have a negative probability density, i.e. ρ < 0.
Technically these problems arise because the Klein-Gordon equation is second order in the
time derivative. Dirac argued that to solve this problem, we need to find an equation that
~ The most general
is linear in ∂/∂t. By Lorentz covariance this also has to be linear in ∇.
form of the equation, for a particle of mass m and wave-function ψ(~x, t), is
∂  
~ + βm ψ(~x, t) .
i ψ(~x, t) = −i~ α·∇ (7.2)
∂t
To be a consistent relativistic equation, ψ also has to obey E 2 = |~p|2 + m2 and be invariant
under Lorentz transformations. So, what are α ~ and β?
Let’s start by requiring that ψ also satisfies the Klein-Gordon equation,
∂ 2 ψ(~x, t) ~ 2 ψ(~x, t) + m2 ψ(~x, t)) .
− 2
= (−∇ (7.3)
∂t
Taking the square of the Dirac equation
  
∂ ∂  
~ + βm −i~

~ + βm ψ(~x, t) ,
i i ψ(~x, t) = −i~α·∇ α·∇ (7.4)
∂t ∂t
= (−iαi ∇i + βm) (−iαj ∇j + βm) ψ(~x, t) , (7.5)
(7.6)
where in the second line we have explicitly included the indices i and j that are being
summed over. This gives
∂ 2 ψ(~x, t)  2 2
− = −αi ∇i + β 2 m2 − (αi αj + αj αi )∇i ∇j (7.7)
∂t2
−i(αi ∇i β + βαj ∇j )m] ψ(~x, t) , (7.8)
where the case that i = j has been separated from the case where i 6= j. We can then
write down the following relations for α
~ and β
β2 =1
αi2 =1
(7.9)
αi αj + αj αi = 2δij
αi β + αj β =0

T. Blake 48 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

An important observation is that αi and β anti-commute i.e. {αi , αj } = 0 if i 6= j and


{αi , β} = 0. They cannot be numbers and instead must be matrices operating on a
multi-component wave-function.
It can be shown that the matrices
• Are Hermitian (αi = a†i );

• Are Traceless, i.e. Tr(αi ) = Tr(β) = 0;

• Have eigenvalues of ±1;

• Have even dimensionality and that the lowest dimension is four.


One popular choice for αi and β is
   
0 σi 12 0
αi = , β= , (7.10)
σi 0 0 −12

where the σi are the Pauli matrices


     
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 = . (7.11)
1 0 i 0 0 −1

The motivation for this choice will become clear later. The Pauli matrices obey the
commutation relation

[σ i , σ j ] = 2iεijk σ k (7.12)

where

 0 if i, j, k repeat

εijk = +1 for cyclic permutations (7.13)

−1 for non-cyclic permutations

If α and β are 4 × 4 matrices, ψ must be a 4-component vector,


 
ψ1
 ψ2
 , ψ † = (ψ1∗ , ψ2∗ , ψ3∗ , ψ4∗ ) .

ψ=  ψ3  (7.14)
ψ4

We can now write the Dirac equation in a more compact form as

(iγ µ ∂µ − m)ψ = 0, (7.15)

with γ µ = (β, βαi ). Explicitly writing out the sum over the space-time indices
∂ψ ∂ψ ∂ψ ∂ψ
iγ 0 + iγ 1 1 + iγ 2 2 + iγ 3 3 − m14 ψ = 0 . (7.16)
∂t ∂x ∂x ∂x

T. Blake 49 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Here, γ µ is not a 4-vector, its components are matrices. Explicitly,


4
" #
X X µ
(iγjk ∂µ − mδjk ) ψk = 0 . (7.17)
k=1 µ

The γ matrices have the following useful properties that we’ll need in our later calculations

γ µ γ ν + γ ν γ µ = 2g µν 1
γ 0† = γ 0
γ 0 γ 0 = +14
(7.18)
γ i† = −γ i
γ i γ i = −14
γ µ† = γ 0 γ µ γ 0

We can also define an adjoint spinor ψ̄ = ψ † γ 0 as the appropriate conjugate of ψ. The


equivalent to the Dirac equation is then

i∂µ ψ̄γ µ + mψ̄ = 0 . (7.19)

To illustrate where this expression comes from, we can take the Hermitian conjugate of
the Dirac equation

−i∂µ ψ † γ µ† − mψ † 1 = 0 . (7.20)

Multiplying from the right by γ 0 gives

−i∂µ ψ † γ µ† γ 0 − mψ † γ 0 = 0 (7.21)
† 0 µ † 0
−i∂µ ψ γ γ − mψ γ = 0
(7.22)

i.e.

i∂µ ψ̄γ µ + mψ̄ = 0 (7.23)

Using ψ and ψ̄ we can form a continuity equation for a current,

∂µ (ψ̄γ µ ψ) = ∂µ j µ = 0 . (7.24)

This solves the problem with the negative probability solutions

ρ = j 0 = ψ̄γ 0 ψ =ψ † γ 0 γ 0 ψ (7.25)
=ψ † ψ > 0 .

T. Blake 50 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

7.2 Free particle solutions


When we derived the Dirac equation we required the wave-function ψ also satisfied the
Klein-Gordon equation. The components ψk of the Dirac spinor, much east satisfying the
Klein-Gordon equation separately, i.e.
 2 
∂ ~ + m ψk = 0 .
2 2
−∇ (7.26)
∂t2
The solutions of this equation will have the general form

ψ = u(p)e−ip·x , (7.27)

where we call u(p) a spinor and e−ip·x is associated with the usual plane wave solution.
Substituting this solution into the Dirac equation yields

(iγ µ ∂µ − m)u(p)e−ip·x = 0 (7.28)

which collapses to

(γ µ pµ − m)u(p) = 0 (7.29)

This is sometimes written with

γ µ pµ = p/ and (p/ − m)u(~p) = 0 . (7.30)

To learn something about the spinor u(p) it is useful to express the γ matrices once again
in terms of α
~ and β,

α · p~ + βm)u(p) = E u(p) .
H u(p) = (~ (7.31)

There are four independent solutions to this equation, two with E > 0 and two with E < 0.
For a particle at rest
 
m12 0
β m u(p) = u(p) = E u(p) (7.32)
0 −m12

which has eigenvalues m, m, −m, −m and eigenvectors


       
1 0 0 0
 0   1   0   0 
  ,   ,   ,   . (7.33)
 0   0   1   0 
0 0 0 1

For p~ 6= 0,
    
m12 ~σ · p~ uA uA
=E (7.34)
~σ · p~ −m12 uB uB

T. Blake 51 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

where u(p) has been divided into two two-component spinor uA and uB . This can be
written more compactly as
(~σ · p~ ) uB = (E − m) uA (7.35)
(~σ · p~ ) uA = (E + m) uB (7.36)
(7.37)
where
 
p3 p1 − ip2
~σ · p~ = . (7.38)
p1 + ip2 −p3
Solutions only exist if
 
−(E − m)1 ~σ · p~
det =0 (7.39)
~σ · p~ −(E + m)1
and we again arrive at E 2 = |~p|2 + m2 , which has positive and negative energy solutions.
The two component spinors are, at this point, arbitrary i.e. the two component spinors
can be written as
uA = a1 χ1 + a2 χ2 (7.40)
uB = b1 χ2 + b2 χ2
with
   
1 1 2 0
χ = , χ = . (7.41)
0 1
For the E > 0 solutions, we can take uA (s) = χ(s) (where s = 1, 2) and
~σ · p~ (s)
uB (s) = χ (7.42)
E+m
The positive-energy four spinors solutions to the Dirac equation are then
χ(s)
 
u(p, s) = N σ ·~
~ p (7.43)
E+m
χ(s)
where N is a normalisation constant. For E < 0, we can instead write uB (s) = χ(s) and
solve for uA such that
 ~σ·~p (s)  σ ·~
~ p
χ(s)
 
E−m
χ − |E|+m
u(p, s + 2) = N =N (7.44)
χ(s) χ(s)
The four solutions are orthogonal, such that u† (r)u(s) = 0 if r 6= s.
For the normalisation, we once again choose N such that we have 2E particles per unit
volume,
 2 !
~σ · p
~
u† (p, s)u(p, s) = |N |2 1 + = 2E . (7.45)
E+m

This sets the normalisation of the spinors to as N = (E + m)1/2 .

T. Blake 52 contact: thomas.blake@cern.ch


Lecture 8

Spin and antiparticles

53
PP2 Relativistic Quantum Mechanics

8.1 Spin
In addition to the positive and negative energy solutions we have an additional two-
fold degeneracy (which gives four eigenvectors/eigenvalues). To understand where this
degeneracy comes from, we need to find out if there is another observable that commutes
with the Hamiltonian.
The Hamiltonian for a free particle is

~ · p~ + βm
H0 = α (8.1)
~ = ~x × p~ commutes with H0 ,
First let’s check to see if angular momentum L
~ = [~
[H0 , L] α · p~ + βm, ~x × p~] (8.2)
α · p~, ~x × p~]
= [~ (8.3)

Using

[AB, C] = A[B, C] + [A, C]B and [pi , pj ] = 0, [xi , pj ] = iδij (8.4)

gives
~ = −i~
[H0 , L] α × p~ (8.5)

i.e. angular momentum is not a good quantum number. The Dirac equation does not
conserve orbital angular momentum and therefore cannot describe a spin-0 particle. Now
lets consider another operator
 
~ = ~
σ 0
Σ (8.6)
0 ~σ
which should be familiar from your quantum mechanics classes.
~ = 2i~
[H0 , Σ] α × p~ (8.7)

The result is equivalent up-to a factor of two and

[H0 , J] ~ + 1 Σ]
~ = [H0 , L ~ =0 (8.8)
2
The operator S~ = 1Σ~ is intrinsic to the particle. For a particle at rest S
~ commutes with
2
H0 and is a good quantum number, e.g. operating through with S3 = 12 Σ3

E > 0 : S3 u(E = +m, 1) = + 12 u(E = +m, 1)


S3 u(E = +m, 2) = − 21 u(E = +m, 2)
(8.9)
E < 0 : S3 u(E = −m, 3) = + 21 u(E = −m, 3)
S3 u(E = −m, 4) = − 21 u(E = −m, 4)

T. Blake 54 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

~ does not commute


which has eigenvalues of ± 12 . We already know that for p~ 6= 0 that S
with H0 . Can we find an operator that does? Let’s try
 
1~ p~ 1~ 1 ~σ · p̂ 0
Σ· = 2 Σ · p̂ = 2 (8.10)
2
|~p| 0 ~σ · p̂

which is the projection of the spin along the particles momentum vector. We refer to this
as the particles Helicity. If we choose p~ = (0, 0, p)
1
E>0: Σ
2 3
u(E > 0, 1) = + 12 u(E > 0, 1)
1
Σ
2 3
u(E > 0, 2) = − 12 u(E > 0, 2)
(8.11)
1
E<0: Σ
2 3
u(E < 0, 3) = + 12 u(E < 0, 3)
1
Σ
2 3
u(E < 0, 4) = − 12 u(E < 0, 4)

Eigenvalues of + 12 have spin aligned with the particles momentum vector. We have found
~ the total angular
that the Dirac equation describes a spin- 12 particle and conserves, J,
momentum.

Figure 8.1: Helicity and handiness of a particle with a momentum vector pointing along the
direction p̂.

T. Blake 55 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Asside:
Helicity is also not a very well defined quantity. We can always boost to the rest-
frame of the particle where the helicity is undefined. A more clearly defined observable
(relativistically covariant observable) is the particles chirality, related to the operator

γ 5 = iγ 0 γ 1 γ 2 γ 3 . (8.12)

In the Pauli-Dirac basis, which we are working with for the γ-matrices, this is
 
5 0 12
γ = . (8.13)
12 0

We can then define left- and right-handed projection operators, which for relativistic
particles project out the left- and right-handed helicities as

PL = 12 (1 − γ 5 )
(8.14)
PR = 12 (1 + γ 5 ) .

T. Blake 56 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

8.2 Antiparticles
It’s interesting to look at how Dirac interpreted the negative energy solutions. He postulated
the existence of a sea of negative energy states and that the vacuum (or ground state of
the system) has all of these negative states full. An additional electron must fill one of the
positive energy states because the Pauli exclusion principal forbids it from occupying a
filled negative energy states. If we promote a negative energy state to a positive one, an
electron-hole pair is created (a positive energy electron and a hole in the negative sea). The
hole appears as a positive energy anti-particle (the positron). This mechanism is essentially
particle and antiparticle pair production. The positron was only observed experimentally
four years later by Anderson in a cloud chamber experiment. Dirac’s mechanism only
works for fermions because bosons do not obey the Pauli exclusion principal.
We have solutions for the Dirac equation that can be written for the E > 0 solutions as

χs
 
−ip·x 1/2
ψ(p+ ) = u(p+ , s)e , u(p+ , s) = (E + m) σ ·~
~ p (8.15)
E+m
χs

and for the E < 0 solutions as


σ ·~
~ p
χs
 
−ip·x 1/2 − |E|+m
ψ(p− ) = u(p− , s + 2)e , u(p− , s + 2) = (E + m) , (8.16)
χs

where

p± = (±E, p~) . (8.17)

If we think of the Feynman-Stückelberg prescription, then we also need to swap the sign
of p~ to interpret the negative energy solutions in terms of anti-particles i.e.
 ~σ·~p s 
1/2 χ
ψ(p)anti = (E + m) E+m
s e+ip·x . (8.18)
χ

We can introduce a second type of spinor for the antiparticles, v(p, s). The v-spinors are
spinors for anti-particles with 4-momentum (E, p~) and we keep the u-spinors for particles
with E > 0. The v-spinors are also orthogonal,

v † (p, r)v(p, s) = 2Eδrs . (8.19)

The two particle solutions of the Dirac equation are

χ1
 
−ip·x 1/2
ψ(p, 1) = u(p, 1)e = (E + m) σ ·~
~ p
E+m
χ1
(8.20)
χ2
 
−ip·x 1/2
ψ(p, 2) = u(p, 2)e = (E + m) σ ·~
~ p .
E+m
χ2

T. Blake 57 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

The two antiparticle solutions of the Dirac equation are


σ ·~
~ p
 
+ip·x 1/2 E+m
χ2
ψ(p, 1) = v(p, 1)e = (E + m) 2
χ
 σ ·~
~ p
 (8.21)
+ip·x 1/2 E+m
χ1
ψ(p, 2) = v(p, 2)e = (E + m) 1 .
χ

Note, the association of v(p, 1) with u(p, 4) and v(p, 2) with u(p, 3). This is due to the
change in the momentum direction and ensures that the indices 1 and 2 in ψ(p, s) refer to
the positive and negative helicity states, respectively.
The current associated to the Dirac equation is

j µ = ψ̄γ µ ψ (8.22)

where ψ̄ is an adjoint spinor defined as ψ̄ = ψ † γ 0 . We also therefore need adjoint u- and


v-spinors. The u-spinors satisfy

(γ µ pµ − m)u(p, s) = (p/ − m)u(p, s) = 0 . (8.23)

The u-spinors for E < 0 with p → −p satisfy

(−γ µ pµ − m)u(−p, s) = 0 (8.24)

or

(γ µ pµ + m)v(p, s) = (p/ + m)v(p, s) = 0 . (8.25)

We also need to define adjoint spinors for our Dirac current j µ = ψ̄γ µ ψ. To obtain an
adjoint spinor we can take the Hermittian conjugate,
 †
(p/ − m)u(p, s) = ū(p, s)(p/ − m) = 0 (8.26)

and similarly

v̄(p, s)(p/ + m) = 0 (8.27)

where

ū(p, s) = u(p, s)† γ 0 and v̄(p, s) = v(p, s)† γ 0 . (8.28)

The combination

u(p, r)† u(p, s) = 2Eδrs , (8.29)

is not Lorentz invariant. However,

ū(p, r)u(p, s) = 2mδrs and v̄(p, r)v(p, s) = −2mδrs . (8.30)

T. Blake 58 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

The orthogonality of the spinors also means that

ū(p, r)v(p, s) = v̄(p, r)u(p, s) = 0 . (8.31)

With the adjoint spinors we have a complete set of states and can write the completeness
relation, summing over the spins
X
u(p, s)ū(p, s) = p/ + m
s=1,2
X (8.32)
v(p, s)v̄(p, s) = p/ − m
s=1,2

This will be useful in our calculations later.

T. Blake 59 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

8.3 Coupling spin- 12 particles


to an EM field
We have already seen how to do this for the Klein-Gordon equation by making the minimal
substitution

i∂ µ → i∂ µ + eAµ , (8.33)

for a particle of charge −e. If we apply this to the Dirac equation

(iγ µ ∂µ − m)ψ = 0 (8.34)

we get

(γ µ (i∂ µ + eAµ ) − m)ψ = 0 (8.35)

There should also be an equivalent expression for a spin- 21 particle with a charge of +e

(γ µ (i∂ µ − eAµ ) − m)ψC = 0 . (8.36)

The form of the equation has to be the same because nature cannot care about how we
define the charges. An obvious question is, what is the relationship between ψ and ψC
(and can we define an operator that transforms between the two)? For the Klein-Gordon
equation, we simply took the complex conjugate. Taking the complex conjugate of the
Dirac equation gives

(γ µ∗ (−i∂µ + eAµ ) − m) ψ ∗ = 0 (8.37)

We can then define,

ψC = Ĉψ ∗ . (8.38)

The operator Ĉ must be a 4 × 4 matrix and satisfy

−Ĉγ µ∗ = γ µ Ĉ , (8.39)

such that

−Ĉ [(γ µ∗ (−i∂µ + eAµ ) − m) ψ ∗ ] = (γ µ (i∂µ − eAµ ) − m) Ĉψ ∗ (8.40)

It must also be Hermitian, i.e. Ĉ † Ĉ = 1. We discussed earlier that the γ-matrices are not
unique and the exact form of Ĉ will depend on the representation of the γ-matrices. In
the Pauli-Dirac representation it is
 
  +1
2 0 iσ2  −1 
Ĉ = iγ = =  (8.41)
−iσ2 0  −1 
+1

T. Blake 60 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

It is instructive to look at what ψC is for our particle wave function, e.g.

ψC (p, 1) = Ĉψ ∗ (p, 1) = iγ 2 u(p, 1)∗ e+ip·x


   
1
 0   e+ip·x
 
= N iγ 2 
 (~ p)∗
σ ·~ 1 
E+m 0
   
σ ·~
~ p 0
E+m
 1   e+ip·x
 
=N 
 0 
1
= u(−p, 4)e−i(−p)·x

This is just

v(p, 1)e+ip·x . (8.42)

The operator Ĉ is the charge conjugation operator. It changes a particle into an anti-
particle and vice versa, flipping the sign of all of the particles charges.

Asside:
The choice of γ matrices is not unique. In this course we are working with the so-called
Pauli-Dirac basis,
   
0 σi 12 0
αi = , β= . (8.43)
σi 0 0 −12

Another popular representation is the Weyl basis


   
−σi 0 0 12
αi = , β= . (8.44)
0 σi 12 0

The Pauli-Dirac basis disgonalises in the particle energy in the non-relativistic limit. This
leads to a particularly simple looking set of eigenvalues and eignevectors for a particle at
rest. The Weyl basis is not diagonalised in the particle energy (the unit matrices in β
appear off-diagonal) and instead diagonalises the particle helicity in the relativistic limit
(such that the unit matrices in γ 5 appear on the diagonal).

T. Blake 61 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

8.4 Gyromagnetic ratio


If we come back to the expression
    
m ~σ · p~ uA uA
=E (8.45)
~σ · p~ −m uB uB
which we can write as a series of coupled equations
~σ · p~ uB = (E − m) uA (8.46)
~σ · p~ uA = (E + m) uB ,
we can rearrange this to write
~σ · p~
uB = uA . (8.47)
E+m
Re-inserting this into the first of the coupled equations gives
1
(~σ · p~ )(~σ · p~ ) uA = (E − m) uA , (8.48)
E+m
or
 
1
E−m− (~σ · p~ )(~σ · p~ ) uA = 0 (8.49)
E+m
In the non-relativisitc limit E ≈ m, such that
1
(~σ · p~)(~σ · p~)uA = 0 (8.50)
2m
Further,
~σ · p~
uB =  uA (8.51)
E+m
We can couple the electron, with charge −e, to an EM field by making the transformation
pµ → pµ + eAµ (8.52)
leading to
 
1 ~ ~
[~σ · (~p + eA)][~σ · (~p + A)] uA = 0 (8.53)
2m
Using p~ = −i∇~ and p~ × ∇~ +A ~ × p~ = i∇
~ × A,~ we get
1 ~ 2 uA + e (∇~ × A)
~ · ~σ uA = 0
(~p + eA) (8.54)
2m 2m
The second term involves a coupling to the magnetic field B ~ =∇~ ×A
~ and the spin of the
~ such that
electron. This is usually written in terms of the magnetic moment as −~µ · B,
e ~ e Σ~
~ =−
µ Σ = −g (8.55)
2m 2m 2
where g is the gyromagnetic ratio of the electron. The prediction that g = 2 is a triumph
of the Dirac equation.

T. Blake 62 contact: thomas.blake@cern.ch


Lecture 9

Coulomb scattering of
charged spin- 12 particles

63
PP2 Relativistic Quantum Mechanics

9.1 Spin- 12 scattering


If you recall for spin-0 particles, the transition amplitude came from first order in pertur-
bation theory and was given by
Z
Tf i = −i φ∗f (x)V (x)φi (x)d4 x (9.1)

For spin- 12 particles we need to make the substitution


Z
Tf i = −i ψf† (x)V (x)ψi (x)d4 x
Z (9.2)
= −i jµf i Aµ d4 x

Once again we need to find the form of the potential, V (x). We can start by writing the
Dirac equation in a form that separates out the Hamiltonian. The Dirac equation is

(iγ µ ∂µ − m)ψ = 0 (9.3)

and we want to write it in a form


∂ψ
H0 ψ = E ψ = i . (9.4)
∂t
We start by expanding the sum over the space-time indices
∂ψ
iγ 0 + γ i ∇i ψ − mψ = 0 , (9.5)
∂t
which can be arranged to give
∂ψ
iγ 0 = (−iγ j ∇j ψ + m)ψ
∂t
∂ψ
iγ 0 γ 0 = (−iγ 0 γ j ∇j ψ + mγ 0 )ψ (9.6)
∂t
∂ψ
i = (−iγ 0 γ j ∇j ψ + mγ 0 )ψ
∂t
If we introduce an EM potential by writing pµ → pµ + eAµ for a particle of charge −e

iγ 0 ∂0 + iγ j ∇j + eγ µ Aµ − m ψ = 0

(9.7)

This can be written as


∂ψ
i = (−iγ 0 γ j ∇j ψ + mγ 0 − eγ 0 γ µ Aµ )ψ (9.8)
∂t

T. Blake 64 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Figure 9.1: Comparison of Feynman rules for spin-0 and spin- 21 particles.

Comparing to the free particle case (with H = H0 + V ), we can make the association
V = −eγ 0 γ µ Aµ . (9.9)
Plugging the potential into the transition amplitude expression gives
Z
Tf i = −i ψf† (x)V (x)ψi (x)d4 x (9.10)
Z
= ie ψf† γ 0 γ µ Aµ ψi d4 x
Z
= ie ψ̄γ µ Aµ ψi d4 x

Again comparing to the spin-0 case, we can identify a transition current


jµf i = −eψ̄f γµ ψi (9.11)
= −eūf γµ ui e+i(pf −pi )·x
For Spin-0 scattering we had
jµf i = −e(pf + pi )µ e+i(pf −pi )·x . (9.12)
Comparing the expressions we see that the vertex factor is ieγ µ and their are now spinors
on the incoming and outgoing legs. The vertex factor is a 4 × 4 matrix.
This brings us to the Feynman rules for spin- 12 particle scattering,

description spin-0 spin- 12

incoming particle 1 u(pi , si )

outgoing particle 1 ū(pf , sf )

incoming antiparticle 1 v̄(pi , si )

outgoing antiparticle 1 v(pf , sf )

vertex factor ie(pi + pf )µ ieγ µ

T. Blake 65 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

We also have internal lines (propagators),


i
spin-0
− m2p2
i(p/ + m)
spin- 12
p2 − m2
ig µν
photon (spin-1) − 2
q
We have exactly the same expression for the spin-0 propagator and the photon propagator.
Does it make sense that these are the same? For a spin-0 particle we derived the photon
propagator from

Aµ = jfµi (9.13)

where jfµi was the transition current associated to the other particle. The dependence on
1/q 2 came from the realisation that
1 −iq·x
∂ µ ∂µ e−iq·x = − e (9.14)
q2
where q = pi − pf .
Finally, we get the expression for the transition amplitude for A + B → C + D,
Z  
(1) 1 µ
Tf i = −i jµ (x) − 2 j(2) (x)d4 x (9.15)
q
 
1
= −i(−eūC γµ uA ) − 2 (−eūD γ µ uB )(2π)4 δ 4 (pA + pB − pC − pD )
q
where q = pA − pC . In terms of the invariant amplitude,

Tf i = −i(2π)4 δ 4 (pA + pB − pC − pD )M (9.16)

where
 
µ −igµν
−iM = (ieūC γ uA ) (ieūD γ ν uB ) (9.17)
q2
All of the dynamics of the system is in the invariant amplitude, M.
At this point it is useful to look in more detail at the structure of the Dirac current,
ūf γ µ ui . To do this we make a so-called Gordon-Decomposition,
1
ūf γ µ ui = ūf ((pf + pi )µ + iσ µν (pf − pi )ν ) ui (9.18)
2m
where

σ µν = 2i [γ µ , γ ν ] = 2i (γ µ γ ν − γ ν γ µ ) (9.19)

T. Blake 66 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

The first term in the decomposition looks identical to what we had in the spin-0 case
for the vertex factor. The second term is a magnetic moment and is related to the
spin of the particle. This is new for the spin- 12 case. If you would like to prove the
Gordon-Decomposition works you can expand the right-hand side and then simplify using

(p/ − m)u(p) = 0 (9.20)


ū(p)(p/ − m) = 0 .

9.2 Interfering diagrams and relative signs


If multiple diagrams need to be included, when calculating an invariant amplitude, it is
also important to account for the relative signs between the different diagrams. Additional
factors of (−1) need to be included in the following scenarios:

1. an anti-fermion line runs continuously from the initial to the final-state;

2. and between diagrams with identical fermions in the final state.

These factors come from the anticommuation properties of the fermionic operators (and
appear when ordering the fermionic fields in a full Quantum Field Theory treatment).

T. Blake 67 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

9.3 Spin- 12 scattering: e−µ− → e−µ−

Figure 9.2: Feynman diagram for spin- 12 e− µ− → e− µ− scattering.

The invariant amplitude for e− µ− → e− µ− scattering is given by


1
−iM = e2 ū(k 0 , s0 )γ µ u(k, s) ū(p0 , r0 )γµ u(p, r) (9.21)
q2
where q = k − k 0 . We want the transition probability for our cross-section calculations
and squaring the invariant amplitude gives
e4
|M|2 = 4
[ū(k 0 , s0 )γ µ u(k, s) ū(p0 , r0 )γµ u(p, r)] (9.22)
q
× [ū(k 0 , s0 )γ ν u(k, s) ū(p0 , r0 )γν u(p, r)]∗

This looks complicated to expand. Don’t worry, it’s not as complicated as it looks. The
structure of ūγ µ u is

(row)(matrix)(column) = (number) (9.23)

and we can write

[āγ ν b]∗ = [āγ ν b]† , (9.24)

More generally we can exploit the structure to split the expression as

[(āγ µ b)(c̄γ ν d)]† = (āγ µ b)† (c̄γ ν d)† (9.25)

T. Blake 68 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Looking a bit closer at the structure we can also write


(āγ µ b)† = b† γ µ† ā† ,
= b† γ µ† γ 0† a ,
(9.26)
= b† (γ 0 γ µ γ 0 )γ 0 a ,
= b† γ 0 γ µ a = b̄γ µ a .
We can now use these blocks to re-write our original expression,
e4
|M|2 = ū(k 0 , s0 )γ µ u(k, s) ū(p0 , r0 )γµ u(p, r) (9.27)
q4
× ū(p, r)γν u(p0 , r0 )ū(k, s)γ ν u(k 0 , s0 )
We can once again exploit the fact that the blocks āγ µ b are numbers to re-arrange this
expression as
e4
|M|2 = ū(k 0 , s0 )γ µ u(k, s) ū(k, s)γ ν u(k 0 , s0 ) (9.28)
q4
× ū(p0 , r0 )γµ u(p, r)ū(p, r)γν u(p0 , r0 )
Why is this a useful thing to do? It identifies a useful feature of the calculation, we can
separate the electron and muon system and write the invariant amplitude
e4 µν (µ)
|M|2 = L L (9.29)
q 4 (e) µν
Tensor associated with the electron vertex is
Lµν 0 0 µ ν 0 0
(e) = ū(k , s )γ u(k, s) ū(k, s)γ u(k , s ) (9.30)
and the muon vertex is
0 0 µ 0 0
L(µ) ν
µν = ū(p , r )γ u(p, r) ū(p, r)γ u(p , r ) (9.31)
In most experiments we are interested in measuring an unpolarised cross-section, i.e. a
cross-section in which we have no information on the spin of the incoming and outgoing
particles and our incoming beams are a equal mixture of the different spin states. To allow
for all possible spin configurations we average over the spins of the incoming particles and
sum over all the possible configurations of the outgoing particles, i.e.
1 X
|M|2 → |M|2 ≡ |M|2 (9.32)
incoming spin-states spin-states

For spin- 12 particles, there are two possible spin-states. We have already seen that summing
of the spins
X
u(p, s)ū(p, s) = (p/ + m) (9.33)
s

T. Blake 69 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

This completeness relation will come in handy for our calculations.


If we look at the structure of the electron and muon tensors, then they have the form

Lµν = āγ µ b b̄γ ν a (9.34)

which in terms of the basic building blocks is

(row)(matrix)(column)(row)(matrix)(column) = (row)(matrix)(column) (9.35)


= (number) (9.36)

Labelling the indices, we can write this as

L = āi Mij aj . (9.37)

Since these are just numbers, they can be re-arranged as

L = aj āi Mij . (9.38)

The product aj āi is another matrix Aji .

Aji Mij = Bjj (9.39)

Summing the indices, this is the trace of a matrix,

L = āM a = Tr(aāM ) (9.40)

T. Blake 70 contact: thomas.blake@cern.ch


Lecture 10

Trace techniques and spin sums

71
PP2 Relativistic Quantum Mechanics

10.1 Trace techniques


As introduced in Sec. 7.1, the Dirac gamma matrices obey the following relations

γ 0 γ 0 = +1
γ i γ i = −1 (10.1)
{γ µ , γ ν } = γ µ γ ν + γ ν γ µ = 2g µν

It is also useful to define

γ 5 = iγ 0 γ 1 γ 2 γ 3 (10.2)

such that
γ 5 γ 5 = 14
γ 5† = γ 5 (10.3)
{γµ , γ5 } = γ µ γ 5 + γ 5 γ µ = 0 .

You will come across the γ 5 matrix again when you deal with interactions involving the
weak interaction which have both vector ūγ µ u and axialvector ūγ µ γ 5 u couplings.
The trace of an n × n matrix is
N
X
Tr(A) = Aii , (10.4)
i=1

where the Aii are the diagonal elements of the matrix. The 4 × 4 unit matrix has a trivial
trace

Tr(1) = 4 . (10.5)

Traces are also the same for cyclic permutations, such that

Tr(ABC) = Tr(BCA) = Tr(CAB) . (10.6)

/ = γ µ aµ we can also write down the following useful rules


Using the notation a

• The trace of an odd number of γ matrices vanishes

a/b) = 4a · b = 4aµ bµ
• Tr(/

a/b/cd/) = 4[(a · b)(c · d) − (a · c)(b · d) + (a · d)(b · c)]


• Tr(/

• Tr(γ 5 ) = 0

• Tr(γ 5 a
//b) = 0
• Tr(γ 5 a
//b/cd/) = 4iεµνλσ aµ bν cλ dσ

T. Blake 72 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Example: To show that the Trace of an odd number of γ matrices vanishes

Tr(γ α γ σ γ ρ . . . γ µ ) = Tr(γ α γ σ γ ρ . . . γ µ γ ν γ 5 γ 5 )
= Tr(γ 5 γ α γ σ γ ρ . . . γ µ γ ν γ 5 )
(10.7)
= Tr(γ 5 γ α γ σ γ ρ . . . γ µ γ 5 γ ν ) × (−1)
= Tr(γ 5 γ 5 γ α γ σ γ ρ . . . γ µ γ 5 γ ν ) × (−1)n

where n is the number of γ matrices and we have used the anti-commutation relation
{γ µ , γ 5 } = 0 and γ 5 γ 5 = 1. If n is odd the trace has to be zero for the left- and right-hand
sides to be equal.

a/b) = 4a · b
Example: To show that Tr(/

a/b) = Tr(/ba
Tr(/ /)
1
a/b + /ba
= 2 Tr(/ /)
= 12 aµ bν Tr(γ µ γ ν + γ ν γ µ )
(10.8)
= 12 aµ bν Tr(2g µν 1)
= g µν aµ bν Tr(1)
= 4g µν aµ bν = 4aν bν

T. Blake 73 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

10.2 Spin- 12 scattering: e−µ− → e−µ−


To progress any further we need to do the spin sums. For the incoming particles we need
to average over the two possible spin configurations. For the outgoing particles we need to
sum over the spin configurations. Taking the expression,

Lµν 0 0 0 0 µ ν
(e) = Tr(u(k , s )ū(k , s )γ u(k, s)ū(k, s)γ ) (10.9)

Applying the spin sums this becomes


1
Lµν /0 µ
/ ν
(e) = Tr((k + m(e) )γ (k + m(e) )γ ) . (10.10)
2
The factor of 12 comes from averaging over the two initial spin states the electron has. The
invariant amplitude can then be written as

1 e4  0 µ ν
  0 
|M|2 = Tr (k/ + m(e) )γ ( /
k + m(e) )γ × Tr ( p
/ + m(µ) )γµ (p/ + m(µ) )γν . (10.11)
4 q4
Both traces have the same structure, so we only need to solve it once,
1  0
Lµν µ ν

(e) = Tr ( /
k + m (e) )γ (k/ + m (e) )γ (10.12)
2
1  0 0
= Tr k/ γ µ k/γ ν + k/ γ µ m(e) γ ν + m(e) γ µ k/γ ν + m2e γ µ γ ν .

2
The easiest term to handle is

Tr (p/γ µ mγ ν ) . (10.13)

This contains an odd number of γ matrices and so the trace is zero. It is also easy to
handle the term
m2
Tr (m2 γ µ γ ν ) = Tr(γ µ γ ν + γ ν γ µ ) (10.14)
2
m2
= Tr(2g µν 1) = 4m2 g µν . (10.15)
2
The remaining term is

Tr(p/0 γ µ p/γ ν ) = Tr(γ σ p0σ γ µ γ ρ pρ γ ν ) (10.16)


= p0σ pρ Tr(γ σ γ µ γ ρ γ ν )
= 4(p0µ pν + p0ν pµ − g µν (p0 · p)) .

Combing the terms

Lµν 0µ ν 0ν µ µν 0 2
(e) = 2(k k + k k − g (k · k − m(e) )) . (10.17)

T. Blake 74 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Finally,

e4 µν (µ)
|M|2 = L Lµν (10.18)
q 4 (e)
4e4
= 4 (k 0µ k ν + k 0ν k µ − g µν (k 0 · k − m2 ))(p0µ pν + p0ν pµ − gµν (p0 · p − m2(µ) )) . (10.19)
q
We now need to multiply all the pieces and keep track of the indices we are summing, e.g.

k µ k 0ν pµ pν = (k · p)(k 0 · p0 ) . (10.20)

We can also use g µν gµν = 4 to simplify the expression. This gives

8e4  0 0 0 0 2 0 2 0 2 2

|M|2 = (k · p)(k · p ) + (k · p )(p · k) − m(e) (p · p ) − m(µ) (k · k ) + 2m(e) m(µ)
q4
(10.21)

If we work in the high energy limit (E  m) we can neglect terms proportional to m or


m2 and write this in the compact form

8e4
|M|2 = [(k · p)(k 0 · p0 ) + (k · p0 )(p0 · k)] , (10.22)
q4
and as before this can be expressed in a frame invariant form in terms of the Mandelstam
variables,

s = +2k · p = +2k 0 · p0 (10.23)


t = −2k · k 0 = −2p · p0
u = −2k · p0 = −2p · k 0

as
s2 + u2
 
4
|M|2 = 2e (10.24)
t2

T. Blake 75 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

10.3 Spin- 12 scattering: e−e+ → µ−µ+


We can obtain the result for e− e+ → µ− µ+ by applying crossing rules (and the Feynman
Stuckelberg interpretation) to the result for e− µ− → e− µ− . This is demonstrated in
Fig. 10.1. The right-hand figure looks a lot like the bottom left-hand figure just viewed
from a different perspective. The crossing tells us we need to exchange

pB ↔ −pC (10.25)

which is equivalent to swapping s ↔ t. The unpolarised result is then


 2 2

4 t + u
|M|2 = 2e (10.26)
s2

Figure 10.1: The relationship between e− e+ → µ− µ+ and e− µ− → e− µ− scattering.

The transition probability can be translated into a differential cross-section for e− e+ →


µ− µ+ scattering using the expression we had from earlier. In the centre of mass system

dσ 1
= |M|2 (10.27)
dΩ 64π 2 s
CM

T. Blake 76 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

which in terms of a scattering angle, θ, can be written as



dσ 1 4 1
 2

= 2e 2
(1 + cos θ) (10.28)
dΩ 64π 2 s
CM
α2
= (1 + cos2 θ)
4s
The total cross section integrating over solid angle (θ and φ) is

4πα2
σ(e− e+ → µ− µ+ ) = (10.29)
3s
This is the first result that we can compare to data (see Fig. 10.2). It agrees quite nicely,
but only in this energy range. At higher energies we can also draw diagrams involving the
Z boson that will have√ a significant impact on the cross-section (there will be a pole-like
enhancement when s = mZ ).

Figure 10.2: Total cross section for e− e+ → µ− µ+ measured at the PETRA accelerator at DESY.
The solid-curve corresponds to our QED expectation.

T. Blake 77 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

10.4 Photons and polarisation vectors


For spin-0 particle scattering we introduced the rule for an incoming/outgoing photon.
We have also seen that Maxwells equations can be written in the compact form

Aµ = j µ with ∂µ Aµ = 0 (10.30)

The requirement ∂µ Aµ = 0 is known as the Lorentz condition. In quantum mechanics, the


wavefunction of the free photon will satisfy

Aµ = 0 (10.31)

i.e. setting the current density to zero. This equation has solutions of the form

Aµ = εµ (~q)e−iq·x (10.32)

Substituting this back into Aµ = 0 gives q 2 = 0, i.e. mγ = 0 (as expected). The
four-vector εµ (~q) deals with the polarisation of the photon. The photon is a spin-1 particle,
which has only two transverse polarisation states and so we expect to be able to cancel
two of the four possible polarisation vectors. In general, a spin-1 particle with non-zero
mass will have three polarisation states. How do we remove two of the polarisation states?
The Lorenz condition ∂µ Aµ gives

qµ ε µ = 0 , (10.33)

reducing the number of independent components to three. We also have freedom in out
choice of gauge, transforming

Aµ → Aµ + ∂µ Λ . (10.34)

Choosing

Λ = iae−iq·x (10.35)

shows that that the underlying physics is unchanged by

ε → ε + aqµ (10.36)

i.e. two polarisation vectors that differ by multiples of qµ describe the same photon. We
can ue this freedom to ensure the time component of the polaristion vanishes, i.e. ε0 = 0.
The Lorentz condition reduces to

~ε · ~q = 0 (10.37)

For a photon travelling along the z-axis, the two polarisation vectors are

ε1 = (1, 0, 0) , ε1 = (0, 1, 0) (10.38)

T. Blake 78 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

It can be shown that the linear combinations


r r
1 1
εR = − (ε1 + iε2 ) , εL = + (ε1 − iε2 ) , (10.39)
2 2
correspond to a photon of helicity of λ = ±1 . The polarisation vectors εL,R are known
as circular polarisation vectors. The circular polarisation vectors obey the completeness
relation
X
(ελ )∗i (ελ )j = δij q̂i q̂j (10.40)
λ

In general a spin-1 particle of helicity λ, mass m and momentum p~ = (0, 0, pz ) can be


described by three polarisation vectors
r
1 1
ελ=±1 = ∓ (0, 1, ±i, 0) , ελ=0 = (|~p|, 0, 0, E) (10.41)
2 m
with the completeness relation
X pµ pν
ελ∗ λ
µ εν = −gµν + (10.42)
λ
m2

T. Blake 79 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Problem set (1)


Special relativity, Lorentz covariance and the Klein-
Gordon equation
1. Show that length
A2 = (A0 )2 − (A1 )2 − (A2 )2 − (A3 )2
is invariant under Lorentz transformation.

[2]

2. Show that gµν g µν = 4.

[2]

3. Using Schrödinger’s equation and the definition of particle density, ρ = ψ ∗ ψ, show


that the system satisfies the continuity equation with a current defined as

~j = 1 ψ ∗ ∇ψ
 
~ − ψ ∇ψ~ ∗ .
2mi

[4]

4. Show that plane waves


~
N e−i(wt−k·~x)
are solutions of the Klein-Gordon equation. Obtain expressions for the energies of
the solutions.

[2]

T. Blake 80 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Problem set (2)


Transition amplitudes and particle scattering
1. Show that solution of the Schrödinger equation
∂ψ
i= [H0 (~x) + κV (t, ~x)] ψ
∂t
written in form of linear superposition of stationary states φm (~x)
X
ψ= am (t)φm (~x)e−iEm t
m

yields a system of differential equations


Z
daf (t) X
i(Ef −Em )t
= −iκ am (t)e φ∗f V (t, ~x)φm (~x)d3~x
dt m
.

[3]

2. Show that the Lorentz invariant phase-space for A + B → C + D scaterring


d3 p~ C d3 p~ D
dQ = (2π)4 δ 4 (pC + pD − pA − pB )
(2π)3 2EC (2π)3 2ED
can be written in polar coordinates as
1 |~p C |
dQ = √ dΩ ,
4π 2 4 s
where dΩ is the element of solid angle and s = (pA + pB )2 . Hence, show that the
differential cross-section for the process is
dσ 1 |~p C |
= |M|2 .
dΩ 64π 2 s |~p A |
[4]

3. In the very high-energy limit (E  m), show that the differential cross section for
spinless electron-muon scattering in the centre of mass system becomes
2
α2 3 + cos θ


= ,
dΩ 4s 1 − cos θ
CM
2
where α = e /4π and θ is the scattering angle.

[3]

T. Blake 81 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Problem set (3)


Dirac equation
1. Using the particle spinors for the positive energy solutions of the Dirac equation,
show that the spinors are orthogonal with a denisty of 2E particles per unit volume,
i.e.
u† (p, r)u(p, s) = 2Eδrs
Further show that
ū(p, r)u(p, s) = 2mδrs
.

[3]

2. Show that spinors u and v satisfy the following relations


X
u(p, s)u(p, s) = γ µ pµ + m,
s
X
v(p, s)v(p, s) = γ µ pµ − m.
s

[3]

3. Consider the operator,  


~ = ~σ 0
Σ
0 ~σ
and show that its commutator with Hamiltonian H0 = α ~ · p~ + βm
is −2i~ ~ H0 ] = −2i~
α × p~ (i.e. [Σ, ~ = +2i~
α × p~ and [H0 , Σ] α × p~).

[4]

For the u-spinors you can take

χ(s)
 
1/2
u(p, s) = (E + m) σ ·~
~ p
E+m
χ(s)

where    
(1) 1 (2) 0
χ = , χ = .
0 1

T. Blake 82 contact: thomas.blake@cern.ch


PP2 Relativistic Quantum Mechanics

Problem set (4)


Spin- 12 particle scattering
1. Calculate the spin averaged amplitude squared (|M|2 ) for the process e+ e− → e+ e−
and in the high energy limit. Express the result in terms of the Mandelstam variables.

Hint: include both diagrams

[10]

T. Blake 83 contact: thomas.blake@cern.ch

You might also like